Polymer Nanosphere Multilayered Organization - ACS Publications

Aug 8, 2015 - AFM images of Z-type single-particle layers of the NVCz:FF10EA:OA = 4:1:1 copolymer on a solid substrate (a) before and (b) after “the...
0 downloads 0 Views 9MB Size
Article pubs.acs.org/Langmuir

Control of Fine Structure in “Polymer Nanosphere Multilayered Organization” and Enhancement of Its Optical Property Atsuhiro Fujimori,*,† Takahiro Kikkawa,† Qi Meng,† and Yuji Shibasaki‡ †

Graduate School of Science and Engineering, Saitama University, 255 Shimo-okubo, Sakura-ku, Saitama 338-8570, Japan Department of Chemistry and Bioengineering, Faculty of Engineering, Iwate University, Ueda 4-3-5, Morioka, Iwate 020-8551, Japan



S Supporting Information *

ABSTRACT: This paper reports on a new functionality exhibited by “polymer nanosphere multilayered organization”, a new type of molecular organization, and the relationship between their structure and function. The polymer nanosphere multilayered organization is a fine structural material formed by the accumulation of single-particle layers of a hydrophobic polymer at the air/water interface; these single-particle layers have uniform height along the c-axis. By employing the “alternate compression−relaxation method”, high-density, lowdefect particle layers are formed with a clear increase in their crystallite sizes. In the case of a ternary comb copolymer containing a carbazole ring, one particle is formed by the assembly of approximately 60 units of collapsed monolayer-like double layers. This structure is stabilized by the formation of side-chain crystals in the interlayer, with oriented π−π stacking of carbazole rings, resulting in enhanced fluorescence emission intensity.



INTRODUCTION In the future, to what degree can molecular technology evolve in the soft interface field? In this study, a newly created molecular organization exhibits functionality through control of the subnanostructure. Noncovalent molecular organizations fabricated by the high control of molecular arrangement have demonstrated excellent functionality, which cannot be achieved in the case of single molecules, through the various intermolecular interactions occurring in that organization and cooperative phenomena based on that organization.1−3 In addition, studies on the morphogenesis of molecular organization are beginning to elucidate the origin of life4,5 as well as induce function enhancement molecular devices.6 Further, an extension of molecular functions may be possible by controlling molecular packing at the subnanometer level associated with mesoscopic morphology for achieving hierarchical arrangement.7,8 Of the several reported examples of “morphological control at the interface”, the following ones are noteworthy; viz., formation of nanofibers,9 nanowires,10,11 nanocoils,12 nanoribbons,13 sea− island structures,14 rods,15 gyroids,16 lamellae,17 and honeycombs.18 In recent years, “polymer nanosphere multilayered organization” has been proposed (Figure S1a).19,20 In these studies, since polymer particles are layered regularly at approximately 5 nm periods, a clear layered period of particles and higher-order reflection are clearly observed by their out-ofplane X-ray diffraction (XRD) profile. The emission intensity of hydrophobic polymers containing a luminescent functional group is enhanced by the formation of multiparticle layers. This enhancement is attributed to the dense packing of lightemitting functional groups in the particles. Furthermore, in this © 2015 American Chemical Society

study, an improvement in the layered regularity and the shift of the long period are reported for the new high-density, low-defect particle films. Here, a monolayer of the hydrophobic polymer is collapsed at the air/water interface, resulting in single particle layer by aggregation of multilayer units as collapsed monolayer. Since the hydrophobic polymer avoids contact with the water surface, the collapsed film is stable (Figure S2). The “polymer nanosphere multilayered organization” is formed by aromatic polyamides, N-vinylcarbazole derivatives, polystyrene, and polyguanamine.19 In the case of polymer monolayers, collapse of monolayer is sometimes referred to as a (three-dimensional) crystal transition.21 Several tens of collapsed monolayer units aggregate, resulting in the formation of nanoparticles at the air/ water interface. Since the height of the particle corresponds to that of a collapsed monolayer, homogeneous height is observed only for a single bilayer or three molecular layers.22,23 On the other hand, since the particle size of the film plane depends on the aggregation number of the aggregated collapsed monolayer units, it is not necessarily uniform. However, clearly observed layered periods and high-order reflection are confirmed by XRD, which was employed to evaluate the height direction for multiparticle layers. This is similar to the formation of colloidal photonic crystals24,25 of single nanometer periods using organic molecules. As a result, this polymer nanosphere multilayered organization is expected to provide a wide range of applications, such as use in optical filters, stained glass, organic optical devices, water-repellent sheets, and reflective material (Figure S1b). Received: March 31, 2015 Revised: August 7, 2015 Published: August 8, 2015 9177

DOI: 10.1021/acs.langmuir.5b01162 Langmuir 2015, 31, 9177−9187

Article

Langmuir

Figure 1. AFM images of Z-type single-particle layers of the NVCz:FF10EA:OA = 4:1:1 copolymer on a solid substrate (a) before and (b) after “the alternate compression−relaxation method”. Out-of-plane XRD profiles of the 20 layers of the NVCz:FF10EA:OA = 4:1:1 copolymer obtained by the LB method (c) before and (d) after “the alternate compression−relaxation method”.

collapse of the polymer monolayer and formation of a singleparticle film with subsequent aggregation cannot be achieved. The control of the particle size and the creation of a high-density, low-defect particle layer,29 which has not been achieved yet, can be achieved by a detailed investigation of conditions for particle formation. Compared to the size of the organized molecular film, that of the integrated particle layer containing constituent units is large, with relatively weak interaction between the structural units. Taking advantage of these characteristics, high-quality particle-aggregated organization is formed. In this study, the formation of particles at the air/water interface was investigated by using a ternary comb copolymer having both hydrocarbon and fluorocarbon side chains. In addition, the formation of high-density, low-defect particle layers was attempted by repeated compression−relaxation.30 Further-

Herein, functionalization via the nanostructure control of internal particles, that is, the elucidation of the relationship between the structure and function, was described. Optical properties were facilitated by the introduction of π-conjugated functional groups in the hydrophobic polymers, which exhibited an interesting change in the particle internal structure. A possible rearrangement occurs through the induction of enhancement of π−π interactions26,27 in the densely aggregated collapsed polymer monolayer. Dense π−π stacking interactions of the emitting moiety also led to the enhancement of its optical property. Unfortunately, at this point, the detailed conditions under which nanoparticles are formed at the air/water interface are still unclear. The most likely factor for inducing their formation is the mobility28 of the polymer on the water surface, without which the 9178

DOI: 10.1021/acs.langmuir.5b01162 Langmuir 2015, 31, 9177−9187

Article

Langmuir

Figure 2. Hysteresis loop of π−A isotherms at 15 °C of the monolayer of the NVCz:FF10EA:OA = 4:1:1 copolymer (solid line: 1st compression; dashed line: relaxation; dotted line: 2nd compression) and schematic of “the alternate compression−relaxation method”. distribution, Mw/Mn ≈ 2.14) estimated by gel-permeation chromatography measurements. According to the theory of the Q−e scheme proposed by Alfrey and Price,34 these ternary comb polymers form alternating copolymers of NVCz and long-chain acrylates. In this case, the e values of NVCz, OA, and FF10EA are −1.40, +1.12, and +0.66, respectively. From these values, it appears that the NVCz:long-chain acrylate copolymers form almost ideal alternating copolymers. Particularly, in this study, the NVCz:OA:FF10EA = 4:1:1 copolymer, which includes high contents of hydrophobic carbazole along with an equal ratio of immiscible amphiphilic monomer units, was mainly used as starting materials for the formation of single-particle layers according to a previous report.19 Formation of Monolayers on the Water Surface and Estimation of Molecular Arrangement. The surface pressure− area (π−A) isotherms were measured using a USI-3-22 film balance (USI Co. Ltd.) at 15 °C. Monolayers and single-particle layers prepared from chloroform solutions (approximately 10−4 M) were formed on distilled water (resistivity is approximately 18 MΩ·cm). First, after the evaporation of chloroform from the solutions for 5 min, the π−A isotherms were recorded at compression speeds of 1.2, 4.8 (standard in this study), and 9.6 cm2 min−1. Next, monolayer films prepared by the Langmuir−Blodgett (LB) method at 10, 15 (standard in this study), and 20 °C were transferred onto glass for XRD studies and mica for atomic force microscopy (AFM) studies.

more, the precise analysis of the internal particle nanostructure was performed to elucidate the relationship between the structure and function contributing to the enhanced lightemitting function. Herein, the formation of two types of sidechain crystals and arrangement of carbazole rings have been confirmed by thin-film XRD and deconvolution of fluorescence spectra.31



EXPERIMENTAL SECTION

Materials. Ternary comb copolymers used in this study were obtained by the copolymerization of N-vinylcarbazole (NVCz) with octadecyl acrylate (OA) and 2-(perfluorodecyl)ethyl acrylate (FF10EA) at various monomer ratios (Figure S3). Copolymerization was conducted in an acetone solution at 50 °C for 48 h using azobis(isobutyronitrile) as the initiator. These monomers and the initiator were purchased from Tokyo Kasei Co. Ltd. and Daikin Fine Chemicals Co. Ltd., respectively, and were used without further purification. The compositions of the ternary comb copolymer were determined by 1H NMR (EX270 NMR, Nihon Denshi Co. Ltd.) spectroscopy. The tacticity of the fluorinated homopolymer obtained by 1 H NMR analysis32 was found to be almost syndiotactic (diad: 58%) according to ref 33. In this study, we synthesized ternary copolymers with molecular weights of approximately 4.43 × 104 (molecular weight 9179

DOI: 10.1021/acs.langmuir.5b01162 Langmuir 2015, 31, 9177−9187

Article

Langmuir

Figure 3. AFM images of the Z-type monolayer or single-particle layers of the NVCz:FF10EA:OA = 4:1:1 copolymer on a solid substrate transferred at (a) 5, (b) 35, and (c) 35 mN m−1 by “the alternate compression−relaxation method”. (d) Schematic of the formation of single-particle layers on the water surface. (e) Size distribution of the single-particle layer. The surface morphologies of the transferred films were observed using a scanning probe microscope (Seiko Instrument, SPA300 with an SPI-3800 probe station) and microfabricated rectangular Si cantilevers with integrated pyramidal tips by applying a constant force of 1.4 N m−1. The long spacing between the layer structures along the c-axis was estimated using an out-of-plane X-ray diffractometer (Rigaku, RintUltima III, Cu Kα radiation, 40 kV, 40 mA) equipped with a graphite monochromator. The in-plane spacing of the two-dimensional lattice of the films was determined by XRD (Bruker AXS, MXP-BX, Cu Kα radiation, 40 kV, 40 mA) equipped with a parabolic graded multilayer mirror.35,36 UV−vis absorption spectra and fluorescence spectra of the LB films on quartz plates were recorded on a spectrophotometer (Jasco V-650) and a spectrofluorometer (Jasco FP-6500; λex = 296 nm), respectively.

density. This low-defect particle layer indicates that the packing of the system resembles hexagonal packing at a plurality of positions. Although the out-of-plane XRD profile of the multiparticle layers, as shown in Figure 1a, exhibits clear reflections (Figure 1c), the differences are obvious by comparison with the high-density film (Figure 1d). Diffraction peaks clearly become sharp. The difference at approximately 200 Å is approximately calculated by the comparison of the diameter of the crystallite size in a direction perpendicular to the (001) plane by using the Scherrer equation.37 Notably, the peak value shifts to low angles. Although a layered organization of particles is composed of exactly the same compound and prepared under the same formation conditions, there is an approximately 10 Å difference in the long period value (as compared to the values of d002 × 2). It is indicated that the distance between the centers of gravity in previous report decreased by the “subduction” of the particles shown in the illustrations (Figure 1c). In the case of high-density, low-defect particle layers, the particles in the upper layers are ordered on top of the particles in the lower layers. Next, the method of forming this high-density particle layer and the reversibility of particle formation are described. Figure 2 shows the reversibility of the π−A curve at 15 °C. At the initial compression stage, a two-dimensional phase transition is observed, i.e., monolayer−single-particle layer transition, which corresponds to the collapse of the monolayer. By hysteresis loop measurements at film expansion after particle film formation, an irreversible π−A curve is obtained. In other words, the singleparticle layer of polymer nanospheres that was once formed at the air/water interface does not reverse back to the monolayer. In addition, during hysteresis loop measurements, the value of the limiting area along the horizontal axis is gradually narrowed. Actually, when this process is performed six times, the fifth and



RESULTS AND DISCUSSION This study investigated the formation of single-particle layers of “polymer nanosphere multilayered organization”, high-density, low-defect particle layers as well as the formation conditions of single-particle layers at the air/water interface, precise analysis of the internal fine structure of particles, and enhancement of optical property. Creation of High-Density, Low-Defect Single-Layer Films of Polymer Nanospheres. Figure 1 shows the formation of high-density, low-defect single-particle layers of the NVCz:OA:FF10EA = 4:1:1 copolymer. The evidence for their formation was supported by AFM, out-of plane XRD, and a change in the fine crystallite size based on the Scherrer equation (the Scherrer constant K = 0.94 was used). In a previous study,19 single-particle layers of the polymer nanospheres formed at the air/water interface have been observed to exhibit many defects and low density (Figure 1a). In contrast, the film (Figure 1b) obtained in this study exhibits low defects and high particle 9180

DOI: 10.1021/acs.langmuir.5b01162 Langmuir 2015, 31, 9177−9187

Article

Langmuir

Figure 4. (a) π−A isotherms of the monolayer at the air/water interface at different surface compression speeds (15 °C, compression speeds: 4.8 cm2/ min, solid line; 9.6 cm2/min, dashed line; and 1.2 cm2/min, dotted line) and AFM images of the Z-type monolayer and single-particle layer on a solid transferred at 35 mN m−1 with compression speeds of (b) 1.2 and (c) 9.6 min/cm2, respectively. (d) Size distribution of the particles in Figure 5c. (e) π− A isotherms of the monolayer on the water surface at different subphase temperatures (compression speed: 4.8 cm2/min; 15 °C, solid line; 20 °C, dashed line; and 10 °C, dotted line) and AFM images of the Z-type monolayer and single-particle layer on a solid transferred at 35 mN m−1 with subphase temperatures of (f) 10 and (g) 20 °C. (h) Size distribution of the particles in Figure 5g.

m−1. At the transfer after the phase transition of the first compression process (35 mN m−1), a low-density single-particle layer with many defects is observed (Figure 3b). After alternate compression−relaxation, there is an obvious improvement in the particle density (Figure 3c). Judging from the values of the limiting area observed in the π−A isotherm and the height data obtained from AFM, the polymer nanospheres are composed of a collapsed unit of polymer monolayer-like bilayer as their structural component (Figure 3d). The average diameter of the nanospheres is 53.7 nm, and their particle distribution is shown in Figure 3e. In a simple calculation, a collapsed film of approximately 60 units is assembled with an aspect ratio of 1:10. That is to say, the actual interfacial nanoparticles are ellipsoid.

sixth curves of the hysteresis loop almost coincide. As a result, the particle layer attains the highest low-defect, high-density state. Hence, we refer to this process as the “alternate compression− relaxation method” for generating a single-particle layer on the water surface. For a system in which the size of components is large and the interaction between the constituent units is relatively weak, a low-defect layer can be effectively formed by this process. Figure 3 provides morphological evidence for the above speculation and internal state of the particles indicated from size information. Figure 3a shows AFM images of the monolayer state before phase transition. Since a phase transition is observed at 10 mN m−1, a monolayer is clearly formed during transfer at 5 mN 9181

DOI: 10.1021/acs.langmuir.5b01162 Langmuir 2015, 31, 9177−9187

Article

Langmuir

Figure 5. Phase diagram of the NVCz:FF10EA:OA = 3:1:1, 4:1:1, and 6:1:1 copolymers at the air/water interface.

temperatures of 10 and 20 °C, such a phase transition is not observed in both isotherms. However, the limiting areas of both isotherms are quite different. The limiting area of the π−A curve at 10 °C is approximately equal to that of the monolayer state at 15 °C. On the other hand, at a subphase temperature of 20 °C, the limiting area is significantly reduced. From the AFM images of the transferred film on the solid substrate, morphologies of the monolayer and single-particle layers were observed at 10 and 20 °C (Figures 4f and 4g, respectively). Also, as compared to the subphase temperature of 15 °C, at 20 °C, there is no significant difference in the particle size distribution (Figure 4h). Figure 5 shows a schematic phase diagram of the NVCz:FF10EA:OA copolymers between monolayer and singleparticle layers by employing surface pressure, subphase temperature, and compression speed as parameters. When the ratio of NVCz to the amphiphilic side-chain unit is low, a very homogeneous monolayer is created. Especially, for the

Study of Formation Conditions of Single-Particle Layers at the Air/Water Interface. Figure 4 shows the dependence of compression speed and subphase temperature on the formation condition of the interfacial particle layer. As indicated by the π−A curves in Figure 4a, the transition from the single-particle layer to the monolayer does not appear at lowspeed compression. Moreover, at a slow compression speed of 1.2 cm2/min, a two-dimensional transition is not observed, and after the transfer at 35 mN m−1, only a homogeneous monolayer is observed (Figure 4b). Even if a compression speed of 9.6 cm2/ min is applied, a large difference in particle size distribution is not observed (Figure 4c,d). Figure 4e shows the temperature dependence of the π−A isotherms of the NVCz:FF10EA:OA = 4:1:1 copolymer.19 As described above, the phase transition from a monolayer to a single-particle layer is confirmed at a subphase temperature of 15 °C and at around 10 mN m−1. However, at subphase 9182

DOI: 10.1021/acs.langmuir.5b01162 Langmuir 2015, 31, 9177−9187

Article

Langmuir

moment of the π-conjugated system, observed by the polarized UV spectrum, and CC double bonds of the carbazole ring, as examined by the polarization IR spectrum, typically show clear incident angle dependence, and the orientation angle of the functional groups calculated from their dichroic ratio is 32°. In other words, functional carbazole units in the polymer nanospheres have a regular oriented structure at certain angles. For estimating the crystallinity and packing system of fluorinated and hydrogenated side chains in the polymer nanospheres, in-plane XRD profile of the multiparticle layers of the NVCz: FF10EA:OA = 4:1:1 copolymer was recorded; the results are shown in Figure 7. Two types of (100) reflection are observed, which correspond to the hexagonal close packing of each side chain. The left signal (dF100 = 5.0 Å) is caused by the short spacing of fluorocarbon chains, and the right signal (dH100 = 4.2 Å) is caused by one of the hydrocarbon chains. Therefore, it seems that both side chains form the phase-separated side-chain crystals (subcell39) inside of nanoparticles each other. Figure 7b also shows an illustration of subcell models of hydrocarbon and fluorocarbon chains. The formation of the side-chain crystals in the interlayer inside the nanoparticles is expected to be a stabilization factor for the bilayer structure with a folded main chain. In this measurement, only the XRD crystalline peak based on the packing of side chains was observed. Considering that the carbazole ring is arranged in the same plane, it is not surprising that the reflection from the π−π stacking of the carbazole ring is confirmed. Spacing of the π−π stacking is commonly known as about 3.85 Å.40 However, as reported in previous paper, polyNVCz homopolymer itself is an amorphous polymer.41 From the previous study,38 although the ordered regularity were able to confirm by the polarized NEXAFS, any diffraction peaks derived from π−π stacking was not obtained by in-plane X-ray diffraction. Although carbazole ring has certain regularity, this ordering might be limiting. Enhancement of Fluorescence Emission Properties. Figure 8 shows the comparison of the normalized intensity of the fluorescence spectra of multilayers and multiparticle layers of the NVCz:FF10EA:OA = 4:1:1 copolymer with the same number of layers. This copolymer emits fluorescence at 370−450 nm. Upon photoexcitation, poly-NVCz exhibits a broad fluorescence

NVCz:FF10EA:OA = 2:1:1 copolymer, a new type of a polymer nanosheet with a homogeneous amorphous surface is formed according to a previous study.38 This observation matches our finding for the NVCz:FF10EA:OA = 4:1:1 copolymer as well as 3:1:1 and 6:1:1 copolymers,19 in that all of the copolymers exhibit the same phase transition behavior. On the other hand, for the NVCz: FF10EA:OA = 1:0:0 copolymer, i.e., a poly-NVCz homopolymer, only a single-particle layer state is formed under all conditions. Under the conditions of high temperature, high pressure, and high compression speed, the transition from the copolymer monolayer to a single-particle layer occurs. In other words, the high mobility of the polymer chains at the air/water interface is found to be suitable. Fine Structure Analysis of Internal Particles. For estimating the internal fine structure of the polymer nanospheres, polarized UV−vis and IR spectra of the multiparticle layers of the NVCz:FF10EA:OA = 4:1:1 copolymer were recorded; the results are shown in Figure 6. The transition

Figure 6. (a) Polarized UV−vis spectra and (b) polarized IR spectra of the LB multiparticle layers (15 °C, 35 mN m−1, and 4.8 cm2/min). (c) Schematic of the orientation of the carbazole ring in the polymer nanospheres.

Figure 7. (a) In-plane XRD profile of the LB multiparticle layers of the NVCz:OA:FF10EA = 4:1:1 copolymer (15 °C, 35 mN m−1, and 4.8 cm2/min). (b) Schematic models of two-dimensional side-chain crystals (subcell) of hydrocarbons and fluorocarbons. 9183

DOI: 10.1021/acs.langmuir.5b01162 Langmuir 2015, 31, 9177−9187

Article

Langmuir

Figure 8. Fluorescence spectra of the LB multiparticle layers and multilayers of the NVCz:OA:FF10EA = 4:1:1 copolymer (15 °C, 35 mN m−1, and 4.8 cm2/min).

spectrum in the violet−blue region, which is attributed to the overlap of emissions from different excimers. The high-energy emission centered at around 370 nm is ascribed to a partially overlapped structure involving only one eclipsed aromatic ring from each carbazole group, while the low-energy emission centered at around 420 nm is proposed to correspond to a “sandwich”-like conformation in which two carbazolyl groups attain an eclipsed or fully overlapped arrangement.13,42 Notably, the normalized emission intensity value of the multiparticle layer is approximately 2 times that of the multilayers. The carbazole ring exhibiting fluorescence with an oriented structure, as discussed in the preceding paragraph, is highly ordered and arranged. A suitable unit distance to the light emission of the carbazole excimer43 is formed in a wide range, and it is considered to efficiently form the excited state. This emission is also confirmed in dilute solutions of the NVCz:FF10EA:OA = 4:1:1 copolymer (Figure 9a). Excimer emission is observed in dilute solutions, mainly originating from the intensity of the emission band, attributed to the partial overlap of the carbazole ring. Of course, in multiparticle layers, light emission by the carbazole excimer is also observed (Figure 9b). However, the intensity of the emission band is stronger, caused by the excimer emission based on overlap of the carbazole

ring. From this result, it is considered that the carbazole rings arranged on the outermost layer inside the particles mainly undergo “sandwich”-like stacking. Compared with the multilayers (Figure 9c), the multiparticle layers exhibit an increase in the ratio of the partial (tilted) stack, which can be observed as a sharp emission band. However, since the ratio of the parallel stacks is not improved, predominant formation of this orientation state is characteristic of the “polymer nanosphere multilayered organization”. Figure 10 shows an illustration of the hierarchical structure of “polymer nanosphere multilayered organization” of the NVCz:FF10EA:OA ternary comb copolymer based on fine structure analysis and light-emitting function measurements. A beautiful hierarchical structure is formed from the functional group level to the layered organization level. Such sophisticated molecular techniques with bottom-up technology would introduce avenues for new science and for the development of new molecular devices.



CONCLUSIONS In this study, fine structure analysis and new functionality produced by “polymer nanosphere multilayered organization”, a new type of molecular organization, are described, as well as the 9184

DOI: 10.1021/acs.langmuir.5b01162 Langmuir 2015, 31, 9177−9187

Article

Langmuir

Figure 9. Fluorescence spectra of (a) chloroform solution (10−4 mol/L), (b) the LB multiparticle layers, and (c) the LB multilayers of the NVCz:OA:FF10EA = 4:1:1 copolymer. (d) Schematic of the stacking of the carbazole rings in the polymer nanospheres.

Figure 10. Schematic models of the hierarchical structure of “polymer nanosphere multilayered organization” of the NVCz:OA:FF10EA = 4:1:1 ternary comb copolymer.

of nanoparticles of the hydrophobic polymer occurs. In the case of a ternary copolymer containing a carbazole ring, one particle is formed by the assembly of approximately 60 units of a collapsed monolayer-like double layer. This structure is stabilized by the formation of side-chain crystals in the interlayer, with oriented π−π stacking of carbazole rings. As a result, fluorescence emission intensity is enhanced, and spectral changes are observed by a change in ring stacking. These results supports the formation

relationship between molecular orientation and functionality. The polymer nanosphere multilayered organization is an elaborate structural material formed by the uniform layering of polymer single-particle layers on the water surface, which have uniform height along the c-axis. Applying the “alternate compression−relaxation process”, high-density, low-defect particle layers are formed, with a clear increase in their crystallite sizes. Under the conditions of high subphase temperature, high pressure, and high compression speed, relatively easy formation 9185

DOI: 10.1021/acs.langmuir.5b01162 Langmuir 2015, 31, 9177−9187

Article

Langmuir

Nanoribbons of Copper Tetracyanoquinodimethane for the Fabrication of Devices and Device Arrays. J. Am. Chem. Soc. 2006, 128, 12917− 12922. (14) Johnston, E.; Bullock, S.; Uilk, J.; Gatenholm, P.; Wynne, K. Networks from α,ω-Dihydroxypoly(dimethylsiloxane) and (Tridecafluoro-1,1,2,2-tetrahydrooctyl) triethoxysilane: Surface Microstructures and Surface Characterization. Macromolecules 1999, 32, 8173−8182. (15) Klok, H.; Langenwalter, J.; Lecommandoux, S. Self-Assembly of Peptide-Based Diblock Oligomers. Macromolecules 2000, 33, 7819− 7826. (16) Liu, F.; Prehm, M.; Zeng, X.; Tschierske, C.; Ungar, G. Skeletal Cubic, Lamellar, and Ribbon Phases of Bundled Thermotropic Bolapolyphiles. J. Am. Chem. Soc. 2014, 136, 6846−6849. (17) Komiyama, H.; Sakai, R.; Hadano, S.; Asaoka, S.; Kamata, K.; Iyoda, T.; Komura, M.; Yamada, T.; Yoshida, H. Enormously Wide Range Cylinder Phase of Liquid Crystalline PEO-b-PMA(Az) Block Copolymer. Macromolecules 2014, 47, 1777−1782. (18) Yabu, H.; Jia, R.; Matsuo, Y.; Ijiro, K.; Yamamoto, S.; Nishino, F.; Takaki, T.; Kuwahara, M.; Shimomura, M. Preparation of Highly Oriented Nano-Pit Arrays by Thermal Shrinking of HoneycombPatterned Polymer Films. Adv. Mater. 2008, 20, 4200−4204. (19) Fujimori, A.; Kaneko, Y.; Kikkawa, T.; Chiba, S.; Shibasaki, Y. Fabrication and Structure of “Polymer Nanosphere Multilayered Organization. J. Colloid Interface Sci. 2014, 418, 338−349. (20) Fujimori, A.; Sato, N.; Chiba, S.; Abe, Y.; Shibasaki, Y. Surface Morphological Changes in Monolayers of Aromatic Polyamides Containing Various N-alkyl Side-Chains. J. Phys. Chem. B 2010, 114, 1822−1835. (21) Joncheray, J. T.; Denoncourt, M. K.; Meier, M.; Schubert, S. U.; Duran, S. R. Two-Dimensional Self-Assembly of Linear Poly(ethylene oxide)-b-poly(ε-caprolactone) Copolymers at the Air-Water Interface. Langmuir 2007, 23, 2423−2429. (22) Kimizuka, N.; Kawasaki, T.; Hirata, K.; Kunitake, T. Supramolecular Membranes. Spontaneous Assembly of Aqueous Bilayer Membrane via Formation of Hydrogen Bonded Pairs of Melamine and Cyanuric Acid Derivatives. J. Am. Chem. Soc. 1998, 120, 4094−4104. (23) Vaknin, D.; Bu, W.; Satija, K. S.; Travesset, A. Ordering by Collapse: Formation of Bilayer and Trilayer Crystals by Folding Langmuir Monolayers. Langmuir 2007, 23, 1888−1897. (24) van Blaaderen, A.; Ruel, R.; Wiltzius, P. Template-Directed Colloidal Crystallization. Nature 1997, 385, 321−324. (25) Rogach, A.; Susha, A.; Caruso, F.; Sukhorukov, G.; Kornowski, A.; Kershaw, S.; Möhwald, H.; Eychmüller, A.; Weller, H. Nano- and Microengineering: Three-Dimensional Colloidal Photonic Crystals Prepared from Submicrometer-Sized Polystyrene Latex Spheres PreCoated with Luminescent Polyelectrolyte/Nanocrystal Shells. Adv. Mater. 2000, 12, 333−337. (26) Leigh, A. D.; Murphy, A.; Smart, P. J.; Deleuze, S. M.; Zerbetto, F. Controlling the Frequency of Macrocyclic Ring Rotation in Benzylic Amide [2]Catenanes. J. Am. Chem. Soc. 1998, 120, 6458−6467. (27) Kobayashi, A.; Dosen, M.; Chang, M.; Nakajima, K.; Noro, S.; Kato, M. Synthesis of Metal-Hydrazone Complexes and Vapochromic Behavior of Their Hydrogen-Bonded Proton-Transfer Assemblies. J. Am. Chem. Soc. 2010, 132, 15286−15298. (28) Akiyama, Y.; Kikuchi, A.; Yamato, M.; Okano, T. Ultrathin Poly(N-isopropylacrylamide) Grafted Layer on Polystyrene Surfaces for Cell Adhesion/Detachment Control. Langmuir 2004, 20, 5506−5511. (29) Muramatsu, M.; Akatsuka, K.; Ebina, Y.; Wang, K.; Sasaki, T.; Ishida, T.; Miyake, K.; Haga, M. Fabrication of Densely Packed Titania Nanosheet Films on Solid Surface by Use of Langmuir-Blodgett Deposition Method without Amphiphilic Additives. Langmuir 2005, 21, 6590−6595. (30) Fujimori, A.; Arai, S.; Kusaka, J.; Kubota, M.; Kurosaka, K. Formation and Structure of Langmuir-Blodgett Films of OrganoModified Aluminosilicate with High Surface Coverage. J. Colloid Interface Sci. 2013, 392, 256−265. (31) Zhao, J.; Jensen, L.; Sung, J.; Zou, S.; Schatz, C. G.; Van Duyne, P. R. Interaction of Plasmon and Molecular Resonances for Rhodamine 6G

of a beautiful hierarchical structure from the functional group level to the layered organization level.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.langmuir.5b01162. Figures S1, S2, and S3 (PDF)



AUTHOR INFORMATION

Corresponding Author

*Tel and Fax: +81-48-858-3503; e-mail: [email protected] (A.F.). Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors significantly appreciate the Ministry of Education, Culture, Sports, Science and Technology (MEXT) for the Grantin-Aid for Scientific Research (C, 25410219 (A.F.)). Finally, I offer my heartfelt condolence to my mentor Professor Kiyoshige Fukuda, Saitama University, who died on July 7, 2015.



REFERENCES

(1) Gaines, Jr., G. L. Insoluble Monolayers at Liquid Gas Interfaces; Wiley: New York, 1966. (2) Petty, M. C. Langmuir-Blodgett Films; Cambridge University Press: New York, 1996. (3) Ulman, A. An Introduction to Ultrathin Organic Films: from Langmuir-Blodgett to Self-assembly; Academic Press: Boston, MA, 1991. (4) Kunitake, T.; Okahata, Y.; Shimomura, M.; Yasunami, S.; Takarabe, K. Formation of Stable Bilayer Assemblies in Water from Single-chain Amphiphiles. Relationship between the Amphiphile Structure and the Aggregate Morphology. J. Am. Chem. Soc. 1981, 103, 5401−5413. (5) Fukuda, K.; Shibasaki, Y.; Nakahara, H.; Liu, M.-h. Spontaneous Formation of Polypeptides in the Interfacial Thin Films of Amphiphilic Amino Acid Esters:Acceleration of the Polycondensation and Control of the Structure of Resultant Polymers. Adv. Colloid Interface Sci. 2000, 87, 113−145. (6) Kuhn, H.; Möbius, D.; Bucher, H. Spectroscopy of Monolayer Assemblies. In Physical Methods of Chemistry; Weissberger, A., Rossiter, B. W., Eds.; Wiley: New York, 1972; Vol. 1, Part IIIB, pp 577−702. (7) Abraham, S.; Vijayaraghavan, K. R.; Das, S. Tuning Microstructures in Organogels: Gelation and Spectroscopic Properties of Mono- and Bis-cholesterol-Linked Diphenylbutadiene Derivatives. Langmuir 2009, 25, 8507−8513. (8) Appel, W.; Portale, G.; Wisse, E.; Dankers, P.; Meijer, E. W. Aggregation of Ureido-Pyrimidinone Supramolecular Thermoplastic Elastomers into Nanofibers: A Kinetic Analysis. Macromolecules 2011, 44, 6776−6784. (9) Wu, C.; Li, M.; Chi, S.; Jun, G. Laser Light Scattering Characterization of a Novel Polymer Nanofiber. Macromolecules 1998, 31, 7553−7554. (10) Huang, M. H.; Mao, S.; Feick, H.; Yan, H.; Wu, Y.; Kind, H.; Weber, E.; Russo, R.; Yang, P. Room-Temperature Ultraviolet Nanowire Nanolasers. Science 2001, 292, 1897−1899. (11) Bumm, L. A.; Arnold, J. J.; Cygan, M. T.; Dunbar, T. D.; Burgin, T. P.; Jones, L.; Allara, D. L.; Tour, J. M.; Weiss, P. S. Are Single Molecular Wires Conducting? Science 1996, 271, 1705−1707. (12) Park, K.; Sung, Y.; Han, S.; Yun, Y.; Hyeon, T. Origin of the Enhanced Catalytic Activity of Carbon Nanocoil-Supported PtRu Alloy Electrocatalysts. J. Phys. Chem. B 2004, 108, 939−944. (13) Liu, Y.; Li, H.; Tu, D.; Ji, Z.; Wang, C.; Tang, Q.; Liu, M.; Hu, W.; Liu, Y.; Zhu, D. Controlling the Growth of Single Crystalline 9186

DOI: 10.1021/acs.langmuir.5b01162 Langmuir 2015, 31, 9177−9187

Article

Langmuir Adsorbed on Silver Nanoparticles. J. Am. Chem. Soc. 2007, 129, 7647− 7656. (32) Yamada, K.; Nakano, T.; Okamoto, Y. Synthesis of Syndiotactic Poly(vinyl alcohol) from Fluorine-Containing Vinyl Esters. Polym. J. 1998, 30, 641−645. (33) Fujimori, A.; Araki, T.; Nakahara, H. Effect of the Atoms in Outermost Layers upon Friction Behaviors of Organized Molecular Films of Comb-Polymers Containing Fluorocarbons. Chem. Lett. 2000, 8, 898−899. (34) Alfrey, T.; Price, C. C. Relative Reactivities in Vinyl Copolymerization. J. Polym. Sci. 1947, 2, 101−106. (35) Fujimori, A.; Sugita, Y.; Nakahara, H.; Ito, E.; Hara, M.; Matsuie, N.; Kanai, K.; Ouchi, Y.; Seki, K. Change of Molecular Packing and Orientation from Monolayer to Multilayers of Hydrogenated and Fluorinated Carboxylate Studied by In-plane X-ray Diffraction together with NEXAFS Spectroscopy at C K-edge. Chem. Phys. Lett. 2004, 387, 345−351. (36) Fujimori, A.; Araki, T.; Nakahara, H.; Ito, E.; Hara, M.; Ishii, H.; Ouchi, Y.; Seki, K. In-Plane X-ray Diffraction and Polarized NEXAFS Spectroscopic Studies on the Organized Molecular Films of Fluorinated Amphiphiles with Vinyl Groups and Their Comb-Polymers. Chem. Phys. Lett. 2001, 349, 6−12. (37) Klug, H. P.; Alexander, L. E. X-ray Diffraction Procedures; John Wiley and Sons: New York, 1974. (38) Fujimori, A.; Hoshizawa, H.; Kobayashi, S.; Sato, N.; Kanai, K.; Ouchi, Y. Highly Ordered and Stable Layered “Polymer Nanosheets” Constructed with Amorphous Side-Chains and π-π Stacking of Functional Groups in Ternary Comb Copolymers. J. Phys. Chem. B 2010, 114, 2100−2110. (39) Vand, V.; Bell, I. P. A Direct Determination of the Crystal Structure of the β Form of Trilaurin. Acta Crystallogr. 1951, 4, 465−469. (40) Higashihara, T.; Ito, S.; Fukuta, S.; Koganezawa, T.; Ueda, M.; Ishizone, T.; Hirao, A. Synthesis and Characterization of ABC-Type Asymmetric Star Polymers Composed of Poly(3-hexylthiophene), Polystyrene, and Poly(2-vinylpyridine) Segments. Macromolecules 2015, 48, 245−255. (41) Hoshizawa, H.; Masuya, R.; Masuko, T.; Fujimori, A. Control of Orientation for Carbazole Group in Comb Copolymers Arranged by Method of Organized Molecular Films. Polym. Adv. Technol. 2007, 18, 353−363. (42) Yokoyama, D.; Park, Y.; Kim, B.; Kim, S.; Pu, Y.; Kido, J.; Park, J. Dual Efficiency Enhancement by Delayed Fluorescence and Dipole Orientation in High-efficiency Fluorescent Organic Light-emitting Diodes. Appl. Phys. Lett. 2011, 99, 123303. (43) Ghosh, D.; Chattopadhyay, N. Characterization of the Excimers of Poly(N-vinylcarbazole) Using TRANES. J. Lumin. 2011, 131, 2207− 2211.

9187

DOI: 10.1021/acs.langmuir.5b01162 Langmuir 2015, 31, 9177−9187